The Origin of the Surprising Stabilities of Highly Charged Self

Publication Date (Web): April 5, 2007 ... Chem. 2007, 46, 9, 3511-3522 .... Ronald Siebert , Andreas Winter , Benjamin Dietzek , Ulrich S. Schubert , ...
0 downloads 0 Views 351KB Size
Inorg. Chem. 2007, 46, 3511−3522

The Origin of the Surprising Stabilities of Highly Charged Self-Assembled Polymetallic Complexes in Solution Gabriel Canard and Claude Piguet* Department of Inorganic, Analytical and Applied Chemistry, UniVersity of GeneVa, 30 quai E. Ansermet, CH-1211 GeneVa 4, Switzerland Received November 9, 2006

The thermodynamics of electrochemical and complexation reactions involving the heterobimetallic triple-stranded helicates [MA(L5)3]n+ (M ) Ru(II), Cr(III) and A ) Ca(II), Lu(III)) reveal that solvation processes mask intramolecular intermetallic repulsions in solution, a phenomenon at the origin of the surprising stabilities of highly charged selfassembled polymetallic complexes in solution. A judicious combination of Born−Haber cycles and the Born equation restores the expected electrostatic trend in the gas phase, in which intermetallic interactions can be simply modeled using a standard Coulombic approach. Semiquantitative estimation and prediction of the contribution of the intermetallic repulsion to the total free energy of the formation of discrete polymetallic assemblies in solution become thus accessible. This point is crucial for programming stable metallosupramolecular architectures in solution.

ation events operating in [Cu3(L1)2]3+ and in[Ag3(L2)2]3+.4-6

Introduction Although the straightforward application of Coulomb eq 1 predicts a significant intermetallic repulsion ∆EMM calcd ) Welec ) 10.4 kJ mol-1 between two adjacent monocharged cations separated by d ) 5.8 Å in the double-stranded helicates [Cu3(L1)2]3+ and [Ag3(L2)2]3+, which disfavors successive metallic complexation, Lehn and co-workers surprisingly reported that the formation of these trinuclear complexes was driven to completion by positive cooperativity (NAv is Avogadro’s number ) 6.023 × 1023 mol-1, zi are the charges of the interacting particles in electrostatic units, e is the elemental charge ) 1.602 × 10-19 C, 0 is the vacuum permittivity constant ) 8.859 × 10-12 C N-1 m-2, r is the relative dielectric permittivity of the medium ) 23 in acetonitrile/chloroform, and d is the intermetallic separation).1,2 A decade later, the thorough investigation of their mechanism of formation3 combined with a careful reexamination of the thermodynamic modeling of self-assembly processes eventually reached the opposite conclusion that negative cooperativity indeed dominates the complex* To whom correspondence should be addressed. E-mail: [email protected]. (1) Lehn, J.-M.; Rigault, A.; Siegel, J.; Harrowfield, J.; Chevrier, B.; Moras, D. Proc. Natl. Acad. Sci. U.S.A. 1987, 84, 2565-2569. (2) (a) Pfeil, A.; Lehn, J.-M. J. Chem. Soc., Chem. Commun. 1992, 838840. (b) Garrett, T. M.; Koert, U.; Lehn, J.-M. J. Phys. Org. Chem. 1992, 5, 529-533. (3) Fatin-Rouge, N.; Blanc, S.; Pfeil, A.; Rigault, A.; Albrecht-Gary, A.M.; Lehn, J.-M. HelV. Chim. Acta 2001, 84, 1694-1711.

10.1021/ic062126o CCC: $37.00 Published on Web 04/05/2007

© 2007 American Chemical Society

∆EMM calcd ) Welec ) -

Navz1z2e2 4π0

∫∞d drr2

(1)

r

However, the best fit of the thermodynamic data leading to [Cu3(L1)2]3+ in acetonitrile/chloroform displayed an attractive microscopic intermetallic interactions ∆ECuCu exp,sol < 0, a mathematical solution discarded at that time because it was counterintuitive and irrelevant when the expected repulsive electrostatic interaction operating between two cations was modeled.5b Although less striking, the fitted value -1 for the intermetallic Eu(III)‚‚‚Eu∆EEuEu exp,sol ) 14 kJ mol (III) interaction, occurring at 9 Å in the polymetallic helicates [Eu3(L3)3]9+, poorly matches the value, ∆EEuEu calcd ) 38 kJ mol-1, computed with eq 1 in a continuous dielectric (r ) 36.1 for acetonitrile).5b Moreover, a reliable prediction of the stability constant for the tetrametallic analogue [Eu4(L4)3]12+ could be only obtained when ∆EEuEu exp,sol ) 14 kJ mol-1 is taken into account.7 This recurrent enigma also (4) Ercolani, G. J. Am. Chem. Soc. 2003, 125, 16097-16103. (5) (a) Hamacek, J.; Borkovec, M.; Piguet, C. Chem.sEur. J. 2005, 11, 5217-5226. (b) Hamacek, J.; Borkovec, M.; Piguet, C. Chem.sEur. J. 2005, 11, 5227-5237. (6) Hamacek, J.; Piguet, C. J. Phys. Chem. B 2006, 110, 7783-7792. (7) Zeckert, K.; Hamacek, J.; Senegas, J.-M.; Dalla-Favera, N.; Floquet, S.; Bernardinelli, G.; Piguet, C. Angew. Chem., Int. Ed. 2005, 44, 7954-7958.

Inorganic Chemistry, Vol. 46, No. 9, 2007

3511

Canard and Piguet Chart 1

arises when a trivalent lanthanide, Ln(III), coordinates to the inert and facial noncovalent tripods [Cr(L5)3]3+. An experimental electrostatic work of only 9 kJ mol-1 is found in acetonitrile for the two triply charged ions at 9.3 Å in [CrLn(L5)3]6+, while eq 1 predicts Welec ) 37 kJ mol-1 (r ) 36.1, dCrLn ) 9.3 Å, Figure 1b).8To design stable molecular or supramolecular assemblies including charged components, this systematic deviation from a simple electrostatic model, which indeed reduces intramolecular intermetallic repulsion, should be addressed. Three possible origins for the rationalization of these phenomena can be envisioned: (i) the breakdown of the dielectric continuum in molecules, which would require the concept of local dielectric constants,9 (ii) specific charge compensation effects in the complexes which affect the Coulombic model, and (iii) specific solvation effects. This contribution thus aims to reconcile experimental intermetallic interactions recorded in solution (∆EMM exp,sol) with a classical Coulombic approach (∆EMM ) derived from calcd eq 1. We focus here on two simple chemical processes, in which the electrostatic contribution to the total free energy change can be easily modeled with eq 1 (Figure 1). First, the Ru(III)/Ru(II)-centered reduction of the triple-stranded helicate [RuIIIA(L5)3](3+z)+ (A ) nothing, noted Ø in the rest of the text, z ) 0; A ) Ca(II), z ) 2; A ) Lu(III), z ) 3) will be considered as a probe for the investigation of a simple charge transfer occurring in the vicinity of a charged but electrochemically inactive spectator cation (Az+, Figure 1a). Second, the complexation process shown in Figure 1b will be compared for two different cationic tripods [Ru(L5)3]2+ and [Cr(L5)3]3+, which differ only in their total charge. Results and Discussions Syntheses and Characterization of the Complexes [Ru(L5)3](CF3SO3)2, [RuCa(L5)3](CF3SO3)4, and [RuLu(L5)3](CF3SO3)5. For bimetallic double- or triple-helical complexes involving heterotopic (i.e., unsymmetrical) ligands, two different isomers exist, depending on the relative arrangement (8) Cantuel, M.; Bernardinelli, G.; Muller, G.; Riehl, J. P.; Piguet, C. Inorg. Chem. 2004, 43, 1840-1849. (9) (a) Richardi, J.; Krienke, H.; Fries, P. H. Chem. Phys. Lett. 1997, 273, 115-121. (b) Sacco, A.; Belorizky, E.; Jeannin, M.; Gorecki, W.; Fries, P. H. J. Phys. II Fr. 1997, 7, 1299-1322.

3512 Inorganic Chemistry, Vol. 46, No. 9, 2007

Figure 1. (a) Schematic thermodynamic Ru-centered reduction of the heterobimetallic triple-stranded helicates [RuIIIA(L5)3](3+z)+ (A ) Ø, z ) 0; A ) Ca(II), z ) 2; A ) Lu, z ) 3). (b) Schematic thermodynamic complexation of trivalent lanthanides, Ln(III), to the facial inert noncovalent receptors [M(L5)3]m+ (M ) Ru(II), m ) 2; M ) Cr(III), m ) 3).

of the strands.10 For [RuA(L5)3](2+z)+ (A ) Ø, z ) 0; A ) Ca(II), z ) 2; A ) Lu(III), z ) 3), the head-to-head-to-head isomer HHH-[RuA(L5)3](2+z)+ results when the three strands

Stabilities of Polymetallic Complexes Scheme 1. Syntheses of Complexes HHH-[RuA(L5)3](CF3SO3)(2+z) (A ) Ø, z ) 0; A ) Ca, z ) 2; A ) Lu, z ) 3) with Numbering Scheme for 1H NMR Experiments

adopt a parallel organization leading to a threefold symmetrical complex, in which Ru(II) is octahedrally coordinated by the three bidentate binding units and Az+ is ninecoordinated by the three tridentate binding units, as exemplified in Figure 1. A reverse arrangement of one strand produces the head-to-head-to-tail C1-symmetrical isomer HHT-[RuA(L5)3](2+z)+. This rather tedious terminology will be used for the discussion of the synthetic strategy to avoid any confusion. However, for the rest of the manuscript, the notation [RuA(L5)3](2+z)+ strictly refers to the axial isomer HHH-[RuA(L5)3](2+z)+. Inspired by a previously published procedure,11 an improved method has been developed for the syntheses of Rucontaining d-f triple-stranded helicates (Scheme 1). HHH[RuLu(L5)3]5+ is thus obtained in a 25% yield by the reaction of Ru(DMSO)6(CF3SO3)2 (1 equiv), L5 (3 equiv), and Lu(CF3SO3)3 (1 equiv) in refluxing ethanol, followed by a careful purification process, which aims to remove the large amount of the undesirable HHT-[Ru(L5)3]2+ and [Ru2(L5)2]4+ side products (Scheme 1 and Figure 2a and b). Upon further treatment of HHH-[RuLu(L5)3]5+ with EDTA4- in acetonitrile, the insoluble lanthanide salt (nBu4N)Lu(EDTA) is separated by filtration, and the resulting noncovalent tripod HHH-[Ru(L5)3]2+ is crystallized as its trifluoromethanesulfonate salt in an 81% yield (Scheme 1 and Figure 2c).8,12 Finally, the recombination of HHH-[Ru(L5)3]2+ with Ca(CF3SO3)2 in acetonitrile provides HHH-[RuCa(L5)3]4+ in a 74% yield (Scheme 1). It is worth noting that HHH-[RuCa(L5)3]4+ cannot be obtained directly by the reaction of Ru(DMSO)6(CF3SO3)2 (1 equiv), L5 (3 equiv), and Ca(CF3(10) Piguet, C.; Bernardinelli, G.; Hopfgartner, G. Chem. ReV. 1997, 97, 2005-2062. (b) Albrecht, M. Chem. ReV. 2001, 101, 3457-3497. (11) Torelli, S.; Delahaye, S.; Hauser, A.; Bernardinelli, G.; Piguet, C. Chem.sEur. J. 2004, 10, 3503-3516. (12) (a) Rigault, S.; Piguet, C.; Bernardinelli, G.; Hopfgartner, G. Angew. Chem., Int. Ed. 1998, 37, 169-172. (b) Rigault, S.; Piguet, C.; Bernardinelli, G.; Hopfgartner, G. J. Chem. Soc., Dalton Trans. 2000, 4587-4600.

Figure 2. 1H NMR spectra of (a) crude HHH-[RuLu(L5)3]5+ and HHT[Ru(L5)3]2+, (b) HHH-[RuLu(L5)3]5+ after purification, (c) HHH-[Ru(L5)3]2+, and (d) HHH-[RuCa(L5)3]4+ (CD3CN, 293 K, numbering in Scheme 1).

SO3)2 (1 equiv), which suggests that, despite the modest overall yield (25%), the templating effect of Lu(III) in the Inorganic Chemistry, Vol. 46, No. 9, 2007

3513

Canard and Piguet 1H

NMR Chemical Shifts (with Respect to TMS) for Ligand L5 and Its Complexes Table 1. HHH-[RuCa(L5)3]4+ in CD3CN at 293 K (Numbering in Scheme 1)

HHH-[RuLu(L5)3]5+,

HHH-[Ru(L5)3]2+, and

bidentate binding site Me1

Me2

H1

H2

H3

H4

H5

H6

H7,8

[RuLu(L5)3]5+

2.42 2.10

4.23 4.43

8.51 7.35

7.64 7.82

8.26 8.28

7.33 7.71

7.20 7.22

7.71 4.87

[Ru(L5)3]2+ [RuCa(L5)3]4+

2.18 2.10

4.11 4.33

7.49 7.34

7.75 7.70

8.27 8.25

7.53 7.53

7.33 7.10

5.71 4.98

4.29 3.55 3.61 3.54 3.48 3.55

L5a

tridentate binding site Me3

H9

H10

H11

H12

H13

H14

H15,16

H17,18

Me4

Me5

[RuLu(L5)3]5+

4.21 4.35

7.69 5.44

7.23 6.99

7.34 7.53

8.38 8.65

7.93 8.28

7.57 7.63

1.13 0.67

4.13 4.31

6.91 5.89

6.61 6.86

7.20 7.35

8.44 8.28

8.04 8.05

7.53 7.49

3.35 3.42, 2.63 3.31 3.00, 2.57

1.30 1.05

[Ru(L5)3]2+ [RuCa(L5)3]4+

3.61 3.42, 2.79 3.54 3.17, 2.80

1.23 1.05

1.05 0.67

L5a

a

Recorded in CDCl3.

preparation of HHH-[RuLu(L5)3]5+ is not negligible. This synthetic process can be followed by 1H NMR spectroscopy for the soluble intermediates and products (Figure 2 and Table 1). Figure 2a shows the crude mixture after the initial assembly step. The upfield shifted signals of the aromatic protons H6 (4.87 ppm) and H9 (5.44 ppm) are diagnostic for the formation of the C3-symmetrical triple-stranded helicate HHH-[RuLu(L5)3]5+ because the wrapping of the three strands puts these protons in the aromatic shielding region of the neighboring ligands in the same complex.11,13 The remaining large number of small peaks can be assigned to the low-symmetrical isomer HHT-[Ru(L5)3]2+, while the insoluble part which was filtered prior to the NMR spectrum being recorded, is tentatively attributed to polymetallic oligomers summarized by the generic label [Ru2(L5)2](CF3SO3)4, in analogy with some closely related observations leading to [Zn2(L5)2](CF3SO3)4.13 The decrease in polarity produced by the addition of CHCl3 to the soluble crude mixture containing HHH-[RuLu(L5)3]5+ and HHT-[Ru(L5)3]2+ induces selective precipitation of the most-charged species, and pure HHH-[RuLu(L5)3](CF3SO3)5 can be separated by filtration. Its redissolution in acetonitrile confirms the existence of the C3-symmetrical HHH-[RuLu(L5)3]5+ as the single species in solution (Figure 2b). The standard decomplexation of Lu(III) with EDTA4- provides the monometallic noncovalent tripod HHH-[Ru(L5)3]2+ (Figure 2c),8,12 in which only H6 (5.71 ppm) is still affected by the wrapping of the strands about Ru(II), whereas H9 (6.91 ppm) appears in the standard domain for an aromatic proton in a diamagnetic complex because the tridentate binding units are noncoordinated (see Figure 1b for a structural scheme). Finally, recombination with Ca2+ restores the closely packed and rigid C3-symmetrical triple-stranded organization of the strands in HHH-[RuCa(L5)3]4+, as measured by the low-field resonances of H6 (4.98 ppm) and H9 (5.89 ppm) and the diastereotopic methylene protons H7,8, H15,16, and H17,18 (Figure 2d).11-13

Elemental analyses confirm the isolation of HHH-[Ru(L5)3](CF3SO3)2‚2.5H2O (1), HHH-[RuCa(L5)3](CF3SO3)4‚ 7H2O (2), and HHH-[RuLu(L5)3](CF3SO3)5‚3.5H2O (3) in the solid state, while the previously reported crystal structure of 3 unambigously established the triple helical character of these complexes, in which the metallic sites are separated by ∼9.0 Å (dRuLu ) 9.08 Å in 3, see Figure 1a).11 Ru-Centered Redox Processes Occurring in [Ru(L5)3](CF3SO3)2, [RuCa(L5)3](CF3SO3)4, and [RuLu(L5)3](CF3SO3)5. The electrochemical Ru(III)/Ru(II) reduction process described in eq 2 and schematically illustrated in Figure 1a has been investigated by cyclic voltammetry (Figure 3 and Table 2). The scan-rate independent 60-80 mV separation between the anodic and cathodic peaks implies the existence of reversible electron-transfer processes, for which the electrochemical potentials 0 ERu(III)A/Ru(II)A can be estimated using standard methods

(13) Piguet, C.; Bu¨nzli, J.-C. G.; Bernardinelli, G.; Hopfgartner, G.; Petoud, S.; Schaad, O. J. Am. Chem. Soc. 1996, 118, 6681-6697.

Figure 3. Cyclic voltamograms of (a) [Ru(L5)3]2+, (b) [RuCa(L5)3]4+, and (c) [RuLu(L5)3]5+ (0.5 mM, CH3CN + 0.1 mM Bu4NPF6, 298 K).

3514 Inorganic Chemistry, Vol. 46, No. 9, 2007

Stabilities of Polymetallic Complexes 0 0 Table 2. Experimental Electrochemical Reduction Potentials (ERu(III)A/Ru(II)A ), Free Energies of Electron Transfer (∆Get,sol ), Hydrodynamic Radii (rh), and Calculated Free Energies of Solvation (∆solvG0) for HHH-[Ru(L5)3]3+/2+, HHH-[RuCa(L5)3]5+/4+, and HHH-[RuLu(L5)3]6+/5+ (Acetonitrile, 298 K)

[Ru(L5)3]3+/2+ [RuCa(L5)3]5+/4+ [RuLu(L5)3]6+/5+

0 a ERu(III)A/Ru(II)A (V)

Eap - Ecpb (mV)

0 c ∆Get,sol (kJ mol-1)

rh,Ru(II)Ad (Å)

0 e ∆solvGRu(III)A (kJ mol-1)

0 e ∆solvGRu(II)A (kJ mol-1)

0.97(1) 0.96(1) 0.98(1)

60(5) 70(7) 80(5)

-94(2) -93(2) -95(2)

7.90(5) 8.10(5) 8.10(5)

-771(4) -2091(12) -3012(19)

-342(3) -1338(7) -2091(12)

a Reduction potential vs standard calomel electrode (SCE) in CH CN + 0.1 mM nBu NPF determined by cyclic voltammetry. b Separation between the 3 4 6 anodic and cathodic peaks. c Calculated with eq 3. d Experimental hydrodynamic radius determined by NMR DOSY (CD3CN, 298 K). e Solvation energies calculated with eq 6 for the Ru(III) and Ru(II) forms, respectively.

(Table 2, columns 2 and 3; we assume that the diffusion coefficients of the analogous Ru(III) and Ru(II) forms are identical within each redox pair).14

Scheme 2. Thermodynamic Born-Haber Cycle for the Ru(III)/Ru(II) Reduction Process Occurring in [RuA(L5)3](3+z)+ (A ) Ø, z ) 0; A ) Ca, z ) 2; A ) Lu, z ) 3)

[RuIIIA(L5)3](3+z)+ + ej f 0 [RuIIA(L5)3](2+z)+ ERu(III)A/Ru(II)A (2)

Surprisingly, the Ru(III)/Ru(II) potentials measured in the complexes [RuIIIA(L5)3](3+z)+ (A ) Ø, z ) 0; A ) Ca(II), z ) 2; A ) Lu(III), z ) 3) and obviously the associated free energies of electron transfer in solution ∆G0et,sol (eq 3, n ) 1 is the number of exchanged electrons and F ) 96490 C mol-1 is the Faraday constant), do not depend on the presence of the spectator cation located at 9.08 Å from the ruthenium center (Table 2, column 4, and Figure 3).11 0 ∆G0et,sol ) -nFERu(III)A/Ru(II)A

(3)

Model and Interpretation of the Electrostatic Intermetallic Communication in [Ru(L5)3](CF3SO3)2, [RuCa(L5)3](CF3SO3)4, and [RuLu(L5)3](CF3SO3)5. Taking into account a continuous relative dielectric permittivity r ) 36.1 in acetonitrile,15 the integration of Coulomb eq 1 gives eq 4, in which Wextra elec is the extra electrostatic work provided by the attachment of an additional electron (z1 ) 1) to Ru(III) in the bimetallic complexes [RuIIIA(L5)3](3+z)+ (i.e., the redox process shown in Figure 1a), because of the additional attractive effect of the spectator cation possessing a z2 charge located at d ) 9.08 Å. Wextra elec ) -

Navz1z2e2 4π0rd

(4)

In absence of a spectator cation, z2 ) 0, and we calIII 3+ culate Wextra elec ([Ru (L5)3] ) ) 0. We can similarly predict III 5+ ([Ru that Wextra Ca(L5) 3] ) ) -8.5 kJ/mol (z2 ) 2) elec extra III 6+ and Welec ([Ru Lu(L5)3] ) ) -12.8 kJ/mol (z2 ) 3), which translate, using eq 3, into cathodic shifts of cath III 3+ III 5+ ∆Ecath shift([Ru (L5)3] ) ) 0 mV, ∆Eshift([Ru Ca(L5)3] ) ) cath III 6+ 89 mV, and ∆Eshift([Ru Lu(L5)3] ) ) 133 mV, respectively. Such large shifts for reversible redox processes cannot escape detection by cyclic voltammetry, and we conclude that this unavoidable electrostatic effect is probably masked in our electrochemical experiments by the change in the solvation energies occurring upon reduction. To get rid of (14) Bard, A. J.; Faulkner, L. R. Electrochemical Methods; John Wiley & Sons: New York, 1980; Vol. 6, pp 213-236. (15) Geary, W. J. Coord. Chem. ReV. 1971, 7, 81-122.

the solvent effects and to restore predictable electrostatic trends, the Born-Haber cycle shown in Scheme 2 has been used to estimate the free energy of the reduction process occurring in the gas phase (∆G0et,gas).16 The latter term is related to the same experimentally accessible process occurring in solution (∆G0et,sol, Table 2, column 4) via the standard free energies of solvation of the complex cations 0 [RuIIIA(L5)3](3+z)+ (∆solvGRu(III)A ) and [RuIIA(L5)3](2+z)+ 0 (∆solvGRu(II)A) according to eq 5. 0 ∆G0et,gas ) ∆G0et, sol + ∆solvGRu(III)A - ∆solvG0Ru(II)A

(5)

To estimate the free energies of solvation required in eq 5, we have used the famous eq 6, proposed by Born in 1920 for a spherical ion in a dielectric continuum.17 The first term of eq 6 corresponds to the notional process of charging a neutral particule having a radius r of the ion concerned in the gas phase (r ) 1), while the second term refers to the same process performed in solvent (r ) 36.1 in acetonitrile). The difference between these two self-energies is interpreted as the Gibbs energy of solvation of the ion in the solvent medium (z is the total charge of the ion in electrostatic units).18 ∆solvG0 ) -

Navz2e2 Navz2e2 + 8π0r 8π0rr

(6)

Except for large-sized monatomic ions, the hypothesis of a spherical ion in the dielectric continuum required in eq 6 is rarely met for real charged systems in chemistry,18 and a number of modifications and adaptations have been proposed to obtain satisfying semiquantitative solvation energies for (16) Chen, H. L.; Ellis, P. E.; Wijesekera, T.; Hagan, T. E.; Groh, S. E.; Lyons, J. E.; Ridge, D. P. J. Am. Chem. Soc. 1994, 116, 1086-1089. (17) Born, M. Z. Phys. 1920, 1, 45-48. (18) Conway, B. E. In Modern Aspects of Electrochemistry; Conway, B. E., White, R. E., Eds.; Kluwer Academic/Plenum Publishers: New York, 2002; pp 295-323.

Inorganic Chemistry, Vol. 46, No. 9, 2007

3515

Canard and Piguet Chart 2

complex ions.19-21As far as the solvation of the cations [RuIIIA(L5)3](3+z)+ and [RuIIA(L5)3](2+z)+ in acetonitrile are concerned, a series of reasonable assumptions can be made for the estimation of their solvation energies. (1) It has been shown by diffusion NMR measurements in polar solutions (r g 25-30) that the formation of ion pairs can be neglected for singly charged cations larger than 4.5-5.0 Å, which thus diffuse independently in solution.22 The recent determination of reliable molecular sizes for the series of double-stranded helicates [Cu(L6)2]+, [Cu2(L7)2]2+, and [Cu3(L8)2]3+ in acetonitrile/chloroforme mixtures by diffusion measurements suggests that these multiply charged cations also diffuse independently in solution (Chart 2),23 an assumption retained for our analogous RuA triple-stranded helicates. (2) The breakdown of the dielectric continuum associated with the polarization of the solvent molecules close to the cation can be neglected when r > 4.5-5.0 Å, a situation encountered for the RuA helicates.19 (3) Small deviations from an ideal sphere for the ions (ellipse or cylinder) has only minor impact on their solvation energies20 and on their autodiffusion coefficients.24 For elliptic or cylindrical length/width ratios p e 3, the approximation of a sphere remains valid;24 a situation which is encountered for [RuIIIA(L5)3](3+z)+ and (19) (20) (21) (22)

Stokes, R. H. J. Phys. Chem. 1964, 86, 979-982. Abe, T. Bull. Chem. Soc. Jpn. 1991, 64, 3035-3038. Kumar, A. J. Phys. Soc. Jpn. 1992, 61, 4247-4250. Martinez-Viviente, E.; Pregosin, P. S.; Vial, L.; Herse, C.; Lacour, J. Chem.sEur. J. 2004, 10, 2912-2918. (23) Allouche, L.; Marquis, A.; Lehn, J.-M. Chem.sEur. J. 2006, 12, 7520-7525. (24) (a) Broersma, S. J. Chem. Phys. 1960, 32, 1626-1631. (b) Doi, M.; Edwards, S. F. The Theory of Polymer Dynamics; Oxford University Press: Oxford, U.K., 1986; Vol. 73, pp 289-303. (c) Galantini, L.; Pavel, N. V. J. Chem. Phys. 2003, 118, 2865-2872. (d) Yamakawa, H.; Tanaka, G. J. Chem. Phys. 2006, 57, 1537-1546.

3516 Inorganic Chemistry, Vol. 46, No. 9, 2007

A(L5)3] because p ≈ 1.6 in the crystal structure of [RuLu(L5)3]5+ (Figure S1a, Supporting Information).11 (4) A simple inspection of eq 6 indicates that, in polar solvents (r g25-30), the second term can be neglected, and ∆solvG0 mainly depends on z2/r in the gas phase.18,21 We conclude that eq 6 can be used for an approximate estimation of the solvation energies of the cations [RuIIIA(L5)3](3+z)+ and [RuIIA(L5)3](2+z)+. However, if the total charges z borne by these cations in the gas phase are obvious, the estimation of their pseudospherical radii is less straightforward. According to Stokes,19 the ionic radius of a monatomic ion in the gasphase significantly deviates from its value calculated by using Shannon’s definition in the solid state.25 A better estimation considers the radius of the noble gas possessing the same electronic configuration, to which a quantum scaling factor is applied for taking into account Slater’s screening effects.19 Following this principle, the size of large cations in the gasphase could be approached by consideration of the volume limited by the Connolly surface constructed around the molecular structure in the crystalline state.26 For the pseudospherical cation [Ru(2,2′-bipyridine)3]2+,27 the Connolly surface, built from the crystal structure, limits a total volume VConnolly ) 450.4 Å3, corresponding to r([Ru(bipy)3]2+ ) ) 3 x3VConnolly/4π ) 4.75 Å if we assume a spherical shape for this cation (Figure S1b, Supporting Information). A related calculation is more debatable for the less-spherical cations [RuIIIA(L5)3](3+z)+ and [RuIIA(L5)3](2+z)+ (Figure S1a, Supporting Information), and we have thus followed Einstein’s original suggestion of using a self-diffusion coefficient in solution for the determination of the real size of molecules.28 For large solute particles (r/rsolvent g 0.5), the StokesEinstein relationship (eq 7) holds,29 and the auto-diffusion coefficient, D, is proportional to r-1, where r is the radius of the moving ion in solution and η is the viscosity of the solvent. Moreover, we assume in our semiquantitative approach that large cations such as [RuIIIA(L5)3](3+z)+ and [RuIIA(L5)3](2+z)+ diffuse in acetonitrile with sufficiently thin solvation spheres that r measured in solution can be taken as a satisfying approximation for the effective radius of the gas-phase cations. Finally, the removal of one electron from a poorly π-bonding Ru-centered orbital on going from [RuIIA(L5)3](2+z)+ to [RuIIIA(L5)3](3+z)+ has a negligible effect on the ionic radius, and we can safely claim that r([RuIIA(L5)3](2+z)+) ≈ r([RuIIIA(L5)3](3+z)+). [RuII

(2+z)+

D)

RT 6πNavηr

(7)

The experimental determination of diffusion coefficients for [RuA(L5)3](2+z)+ (A ) Ø, z ) 0; A ) Ca(II), z ) 2; A ) Lu(III), z ) 3) using 1H diffusional ordered spectroscopy (25) Shannon, R. D. Acta Crystallogr. 1976, A32, 751-767. (26) (a) Connolly, M. L. Science 1983, 221, 709-713. (b) Connolly, M. l. J. Appl. Crystallogr. 1983, 16, 548-553. (27) Harrowfield, J. M.; Sobolev, A. N. Aust. J. Chem. 1994, 47, 763767. (28) (a) Einstein, A. Ann. Phys. 1906, 19, 289-306. (b) (a) Einstein, A. Ann. Phys. 1906, 19, 371-381. (29) Sharma, M.; Yashonath, S. J. Phys. Chem. B 2006, 110, 17207-17211.

Stabilities of Polymetallic Complexes 0 Table 3. Gibbs Free Energy (∆Get,gas ) for the Ru(III)/Ru(II) Reduction extra ) and Process Occurring in the Gas Phase and Experimental (Welec extra Calculated (Welec,calcd ) Extra Electrostatic Work for HHH-[Ru(L5)3]3+/2+, HHH-[RuCa(L5)3]5+/4+, and HHH-[RuLu(L5)3]6+/5+ (Acetonitrile, 298 K)

[Ru(L5)3]3+/2+ [RuCa(L5)3]5+/4+ [RuLu(L5)3]6+/5+

0 a ∆Get,gas (kJ mol-1)

extrab Welec (kJ mol-1)

extra c Welec,calcd (kJ mol-1)

-523 -846 -1016

0 -323 -493

0 -306 -459

a Calculated with eq 5. b Calculated with eq 8. c Calculated with eq 4 (r ) 1, d ) 9.08 Å).

(DOSY) in CD3CN at 298 K gives D([Ru(bipy)3]2+) ) 13.6(4) × 10-10 m2 s-1, D([Ru(L5)3]2+) ) 8.2(1) ×10-10 m2 s-1, and D([RuCa(L5)3]4+) ) D([RuLu(L5)3]5+) ) 8.0(1) × 10-10 m2 s-1. The first value can be compared with the value, D([Cu(L6)2]+) ) 12.5 × 10-10 m2 s-1, reported for a pseudotetrahedral Cu(I) cation with the bulky ligand L6,23 while the autodiffusion coefficients for the triple-stranded ruthenium helicates are similar to that, D([Cu2(L7)2]2+) ) 8.8 × 10-10 m2 s-1, measured for an analogous binuclear double-stranded helicate in the same experimental conditions.23 With r([Ru(bipy)3]2+) ) 4.75 Å as a reference, eq 7 gives r([Ru(L5)3]2+) ) 7.9 Å and r([RuCa(L5)3]4+) ) r([RuLu(L5)3]5+) ) 8.1 Å (Table 2, column 5), in fair agreement with value, r([Cu2(L7)2]2+) ) 7.2 Å, calculated for the related double-stranded helicate in which the two cations are only separated by 5.6 Å.23 The introduction of these cationic radii in eq 6 eventually leads to the solvation energies ∆solvG0 collected in Table 2 (columns 6 and 7). The introduction of ∆solvG0 (Table 2, columns 6 and 7) and ∆G0et,sol (Table 2, column 4) into eq 5 allows the estimation of the Gibbs energy change accompanying the Ru(III)/Ru(II) reduction process performed in the gas phase (∆G0et,gas, Table 3). The latter free energy term, ∆G0et,gas, can be partitioned between (i) the electrostatic work and the electronic and structural reorganizations accompanying the reduction of the Ru(III)N6 site, λRu(III)/Ru(II), a contribution which is reasonably assumed to be identical for the three complexes [Ru(L5)3]3+, [RuCa(L5)3]5+ and [RuLu(L5)3]6+, and (ii) the extra electrostatic work, Wextra elec , generated by the approach of one electron with respect to the spectator cation located close to the ruthenium center in [RuCa(L5)3]5+ and [RuLu(L5)3]6+ (d ) 9.08 Å, eq 8).11 ∆Get,gas0 ) λRu(III)/Ru(II) + Wextra elec

(8)

For [Ru(L5)3]3+, Wextra elec ) 0 and thus λRu(III)/Ru(II) ) 0 3+ ∆Get,gas([Ru(L5)3] ) ) -523 kJ mol-1, from which -1 and Wextra(RuLu)) -493 kJ Wextra elec (RuCa) ) -323 kJ mol elec mol-1 can be easily calculated with eq 8 and the ∆G0et,gas values collected in Table 3. Interestingly, the expected extra electrostatic trend |Wextra elec (RuCa)| < |Welec (RuLu)| is qualitatively restored in the gas phase, and these values can be satisfactorially compared with Wextra elec,calcd(RuCa)) -306 kJ (RuLu)) -459 kJ mol-1 calculated mol-1 and Wextra elec,calcd with Coulomb eq 4 in the gas phase (r ) 1, Table 3). The minor 8% discrepancy between experimental and

Figure 4. (a) Variation of absorption spectra observed for the spectrophotometric titration of [Ru(L5)3]2+ (10-4 mM in acetonitrile) with La(CF3SO3)3·3H2O at 298 K (La/[Ru(L5)3]2+ ) 0.1-2.3). (b) Corresponding variation of observed molar extinctions at four different wavelengths.

predicted data in the gas phase is remarkable according to the rough semiquantitative approach used for the estimation of the solvation energies. We thus conclude that the intermetallic communication in [RuIIIA(L5)3](3+z)+ and in [RuIIA(L5)3](2+z)+ has indeed a simple electrostatic origin in the gas phase, but it is masked in solution by opposite solvation effects which favor the stabilization of highly charged ions. Complexation Processes Occurring in [RuLn(L5)3](CF3SO3)5 and [CrLn(L5)3](CF3SO3)6. Spectrophotometric titrations of [Ru(L5)3]2+ (10-4 M, CH3CN, 298 K) with Ln(CF3SO3)3‚xH2O (Ln ) La, Eu, Gd, Er, Lu; x ) 1-4) show a smooth evolution of the absorption spectra resulting from the red-shift of the ligand-centered π f π* transitions occurring upon complexation of the tridentate binding unit to Ln(III), as previously established for the same complexation reaction performed with [Cr(L5)3]3+ instead of [Ru(L5)3]2+ (Figure 4).8 A single end point is systematically observed for Ln/[Ru(L5)3]2+ ) 1.0, together with two isosbestic points at 273 and 328 nm, implying the existence of only two species [Ru(L5)3]2+ and [RuLn(L5)3]5+ absorbing in the near UV (Figure 1b). The spectrophotometric data can be satisfyingly fitted with nonlinear least-squares technique to equilibrium 9.30 The formation constants, log(βRuLn 11 ) ) 5.2-5.4, do not vary significantly along the lanthanide series within experi(30) (a) Gampp, H.; Maeder, M.; Meyer, C. J.; Zuberbu¨hler, A. Talanta 1986, 33, 943-951. (b) Gampp, H.; Maeder, M.; Meyer, C. J.; Zuberbu¨hler, A. Talanta 1985, 23, 1133-1139.

Inorganic Chemistry, Vol. 46, No. 9, 2007

3517

Canard and Piguet [Ru(L5)3]2+ + Ln3+ h [RuLn(L5)3]5+ log (βRuLn 11 ) (9) mental errors and point to negligible size-discriminating effects (Table 4). Moreover, the log(βRuLn 11 ) values are similar to the log(βCrLn ) values previously reported for the 11 same complexation processes recorded under the same conditions, but in which the doubly charged tripod [Ru(L5)3]2+ is replaced with the triply charged analogue [Cr(L5)3]3+ (Table 4).8 Model and Interpretation of the Electrostatic Intermetallic Communication in [RuLn(L5)3](CF3SO3)5 and [CrLn(L5)3](CF3SO3)6. The experimental formation constants determined for the above-mentioned complexation processes again suggest that the expected increased electrostatic repulsion operating in [CrLn(L5)3]6+ is overcome in solution by the reorganization of the solvent around the cation. This can be modeled by the Born-Haber cycle depicted in Scheme 3, whereby the free energy of complexation in the gas phase (∆G0comp,gas(MLn)) is related to that in solution (∆G0comp,sol(MLn)) by the balance of the various solvation energies (eqs 10 and 11). Moreover, ∆G0comp,gas(MLn) can be partitioned between (i) the sum of the free energies of binding and of reorganization of [M(L5)3]m+ required to accommodate Ln(III), δLn,gas, a contribution which is reasonably assumed to be identical for the two noncovalent tridodal receptors [Ru(L5)3]2+ and [Cr(L5)3]3+, and (ii) the extra electrostatic work Wextra elec,MLn generated by the approach of the tricationic lanthanide Ln3+ from the Mm+ cations in [RuLn(L5)3]5+ (m ) 2, d ) 9.08 Å)11 and [CrLn(L5)3]6+ (m ) 3, d ) 9.35 Å).31 ∆G0comp,gas(RuLn) ) ∆G0comp,sol(RuLn) + ∆solvG0(Ru) + extra (10) ∆solvG0(Ln) - ∆solvG0(RuLn) ) δLn + Welec,RuLn

∆G0comp,gas(CrLn) ) ∆G0comp,sol(CrLn) + ∆solvG0(Cr) + extra (11) ∆solvG0(Ln) - ∆solvG0(CrLn) ) δLn + Welec,CrLn

The difference between eqs 11 and 10 leads to eq 12, in which the variation of the free energy of complexation in the gas phase (∆G0comp,gas(CrLn) - ∆G0comp,gas(RuLn)) indeed corresponds to the difference of the electrostatic works required for bringing Ln3+ close to Cr3+ and to Ru2+, extra extra respectively, in [MLn(L5)3](m+3)+ (Welec,CrLn - Welec,RuLn ). extra ∆G0comp,gas(CrLn) - ∆G0comp,gas(RuLn) ) Welec,CrLn extra ) ∆G0comp,sol(CrLn) - ∆G0comp,sol(RuLn) + Welec,RuLn

∆solvG0(Cr) - ∆solvG0(Ru) + ∆solvG0(RuLn) ∆solvG0(CrLn) (12) The introduction of the solvation energies collected in Table 2 (we assume ∆solvG0([Cr(L5)3]3+) ) ∆solvG0([Ru(L5)3]3+) ) -771 kJ mol-1 and ∆solvG0([CrLu(L5)3]6+) (31) Cantuel, M.; Bernardinelli, G.; Imbert, D.; Bu¨nzli, J.-C. G.; Hopfgartner, G.; Piguet, C. J. Chem. Soc., Dalton Trans. 2002, 19291940.

3518 Inorganic Chemistry, Vol. 46, No. 9, 2007

log(βMLn 11 )

Table 4. Formation Constants for the Complexes [MLn(L5)3](m+3)+ According to Equilibrium 9 (M ) Ru, m ) 2; M ) Cr, m ) 3; Ln ) La, Nd, Eu, Gd, Er, Lu; Acetonitrile, 298 K) Ln(III)

log(βRuLn 11 )

0 (RuLn) ∆Gcomp,sol (kJ mol-1)

La Nd Eu Gd Er Lu

5.4(2)

-31(1)

5.2(2) 5.4(2) 5.4(2) 5.2(2)

-30(1) -31(1) -31(1) -30(1)

a

a log(βCrLn 11 )

0 ∆Gcomp,sol (CrLn)a (kJ mol-1)

5.9(3) 5.4(3)

-34(2) -31(2)

5.4(3) 5.3(3) 5.3(3)

-31(2) -30(2) -30(2)

Formation constants taken from ref 8.

Scheme 3. Thermodynamic Born-Haber Cycle for the Complexation of Ln(III) to [M(L5)3]m+ to Give [MLn(L5)3](m+3)+ (M ) Ru, m ) 2; M ) Cr, m ) 3; Ln ) La-Lu)

) ∆solvG0([RuLu(L5)3]6+) ) -3012 kJ mol-1), together with the free energy of complexation for Ln ) Lu (∆G0comp,sol(MLu) ) -30 kJ mol-1, Table 4) into eq 12 extra extra - Welec,RuLu ) 492 kJ mol-1. This experigives Welec,CrLu mental value can be fairly compared with the difference of the electrostatic works, each being calculated with Coulomb eq 4 (z1 ) z2 ) 3 for [CrLu(L5)3]6+; z1 ) 2 and z2 ) 3 for [RuLu(L5)3]5+), and eventually combined in eq 13 to give extra extra Welec,CrLu,calcd - Welec,RuLu,calcd ) 419 kJ mol-1 (dRuLu ) 9.08 Å11 and dCrLu ) 9.35 Å).31 extra extra Welec,CrLn,calcd - Welec,RuLn,calcd )

[

]

3e2Nav 2 3 (13) 4π0 dRuLn dCrLn The slightly larger discrepancy (15%) between the Coulomb prediction in the gas phase (419 kJ mol-1) and the experimental data extracted from the complexation processes described in eq 9 (492 kJ mol-1), compared with the only 8% discrepancy obtained when modeling the reduction process described in eq 5 may have two origins. (1) Four approximate solvation energies (semiquantitative approach) are required to transform ∆G0comp,sol(CrLn) ∆G0comp,sol(RuLn) into ∆G0comp,gas(CrLn) - ∆G0comp,gas(RuLn) (eq 12), while only two are necessary for correlating ∆G0et,sol and ∆G0et,gas (eq 5). (2) Some polarization of the coordination bonds involving the d-block ions Ru(II) and, especially, Cr(III) may reduce their effective charge in the bimetallic complexes [MLn(L5)3](m+z)+ and thus affects the predictions obtained with eq 13. This limitation is removed when the extra electrostatic work associated with the reduction process occurring in [RuLn(L5)3](3+z)+ is considered because the hard 4s- or 4f-block spectator cations are poorly polarizable.

Stabilities of Polymetallic Complexes

The solvation energies of each microspecies t-[Eu(L3)3]3+, ct-[Eu2(L3)3]6+, and [Eu3(L3)3]9+ can be estimated using Born eq 6, thus leading to eqs 19-21, where r ) 36.1 for acetonitrile and r ) r(t-[Eu(L3)3]3+) ) 1.25r([Ru(L5)3]2+) ) 10 Å because the trinuclear triple-stranded helicate [Eu3(L3)3]9+ possesses three equally spaced metals, while the bimetallic analogue [RuLn(L5)3]5+ displays only two metals separated by the same distance. The 1.25 ratio corresponds to the average value established by NMR diffusion measurements recorded on going from the binuclear double-stranded helicate [Cu2(L7)2]2+ to the trinuclear analogue [Cu3(L8)2]3+ (Chart 2) and for closely related systems.23 Finally, the stepwise rigidification of the triple helix associated with the successive reaction of Eu(III) with t-[Eu(L3)3]3+ to give ct[Eu2(L3)3]6+ and finally [Eu3(L3)3]9+ (Figure 5) produces a stepwise increase of ∼10% in the ionic radii (eqs 19-21) as observed when Ln(III) cooordinates to [Ru(L5)3]2+ to give [RuLn(L5)3]5+. Figure 5. Thermodynamic Born-Haber cycles for the successive complexation of Eu(III) to (a) t-[Eu(L3)3]3+ to give ct-[Eu2(L3)3]6+ and (b) ct-[Eu2(L3)3]6+ to give [Eu3(L3)3]9+.

Application of the Solvation Model for Unravelling Electrostatic Intermetallic Communication in Solution. -1 The incomprehensible small value ∆EEuEu exp,sol ) 14 kJ mol characterizing the intermetallic Eu(III)‚‚‚Eu(III) interaction occurring at 9 Å in the self-assembly of the trinuclear triplestranded polymetallic helicate [Eu3(L3)3]9+ in acetonitrile,5b which significantly contributes to its surprising stability, can be now re-examined in light of the combination of electrostatic and solvation effects previously developed for the rationalization of redox and complexation processes occurring for d-f bimetallic helicates in solution. Let us consider the formation of the three microspecies t-[Eu(L3)3]3+, ct[Eu2(L3)3]6+, and [Eu3(L3)3]9+ characterized by equilibria 14-16 (c ) central and t ) terminal correspond to the location of the europium atoms in the central N9 and terminal N6O3 binding sites, respectively, provided by the virtually preorganized receptor (L3)3, see Figure 5 for the schematic structures of the complexes).5b 3L3 + Eu3+ h t-[Eu(L3)3]3+ log(βt13) ) 16.2

(14)

3L3 + 2Eu3+ h ct-[Eu2(L3)3]6+ log(βct23) ) 26.1 (15) 3L3 + 3Eu3+ h [Eu3(L3)3]9+ log(βtct 33) ) 34.8

(16)

The associated successive constants K2 and K3, corresponding to equilibria 17 and 18, allow the calculation of the free energies for the successive complexation of Eu(III) to t-[Eu(L3)3]3+ (∆GK2,sol ) -56.5 kJ mol-1) and to ct-[Eu2(L3)3]6+ (∆GK3,sol ) -49.6 kJ mol-1) in solution (Figure 5).5b t-[Eu(L3)3]3+ + Eu3+ h ct-[Eu2(L3)3]6+ log(K2) )

log(βct23)

-

log(βt13)

) 9.9 (17)

ct-[Eu2(L3)3]6+ + Eu3+ h [Eu3(L3)3]9+ ct log(K3) ) log(βtct 33) - log(β23) ) 8.7 (18)

∆solvG0(Eu(L3)3) ) -

∆solvG0(Eu2(L3)3) ) -

( )

9Nave2 1 1) 8π0r r

36Nave2

-607 kJ mol-1 (19)

( )

8π0(1.1r)

1-

1 ) r

-2208 kJ mol-1 (20) ∆solvG (Eu3(L3)3 )) 0

81Nave2

( )

8π0(1.2r)

1-

1 ) r

-4554 kJ mol-1 (21) According to the thermodynamic Born-Haber cycles depicted in Figure 5, the Gibbs free energy for the successive complexation of Eu(III) to t-[Eu(L3)3]3+ (equilibria 17 and 18) in the gas phase can be estimated with eqs 22 and 23. Since the affinity of Eu(III) for the N9 and N6O3 coordination sites is known to be very similar,32,33The gas-phase free energies of successive complexation can be partitioned between (i) the intermolecular free energy of connection of the entering Eu(III) in the nine-coordinate site δcomp, a contribution which is reasonably assumed to be identical for the two successive coordination processes, (ii) the extra electrostatic work Wextra elec,adjacent for the approach of two adjacent Eu3+ in ct-[Eu2(L3)3]6+ and in [Eu3(L3)3]6+ (d ) 9 extra Å)32 and (iii) the extra electrostatic work Welec,distal for the approach of two distal Eu3+ in [Eu3(L3)3]6+ (d ) 18 Å, eqs 22 and 23).32 The difference in free energy of these two successive complexation processes is given in eq 24. (32) (a) Floquet, S.; Ouali, N.; Bocquet, B.; Bernardinelli, G.; Imbert, D.; Bu¨nzli, J.-C. G.; Hopfgartner, G.; Piguet, C. Chem.sEur. J. 2003, 9, 1860-1875. (b) Floquet, S.; Borkovec, M.; Bernardinelli, G.; Pinto, A.; Leuthold, L.-A.; Hopfgartner, G.; Imbert, D.; Bu¨nzli, J.-C. G.; Piguet, C. Chem.sEur. J. 2004, 10, 1091-1105. (33) (a) Piguet, C.; Borkovec, M.; Hamacek, J.; Zeckert, K. Coord. Chem. ReV. 2005, 249, 705-726. (b) Hamacek, J.; Borkovec, M.; Piguet, C. Dalton Trans. 2006, 1473-1490.

Inorganic Chemistry, Vol. 46, No. 9, 2007

3519

Canard and Piguet 0 0 ∆GK2,gas ) ∆GK2,sol + ∆solvG0(Eu(L3)3) + ∆solvG0(Eu) -

∆solvG0(Eu2(L3)3) ) δcomp + Wextra elec,adjacent (22) 0 0 ) ∆GK3,sol + ∆solvG0(Eu2(L3)3) + ∆solvG0(Eu) ∆GK3,gas extra ∆solvG0(Eu3(L3)3) ) δcomp + Wextra elec,adjacent + Welec,distal (23) 0 0 0 0 - ∆GK2,gas ) ∆GK3,sol - ∆GK2,sol + ∆GK3,gas

2∆solvG0(Eu2(L3)3) - ∆solvG0(Eu(L3)3) extra (24) ∆solvG0(Eu3(L3)3) ) Welec,distal

The introduction of eqs 19-21, together with the experimental values ∆GK2,sol ) -56.5 kJ mol-1 and ∆GK3,sol ) 0 0 -49.6 kJ mol-1, into eq 24 leads to ∆GK3,gas - ∆GK2,gas ) -1, a value which can be compared with the 752 kJ mol extra Welec,distal ) 694 kJ mol-1 value calculated with the simple Coulomb eq 4 in the gas phase (z1 ) z2 ) 3, r ) 1, d ) 18 Å).32 Again, the straightforward electrostatic model is satisfying within 10% for the rationalization of intermetallic interactions in molecules, according that semiquantitative solvation energies are explicitely considered. Finally, it is interesting to estimate the standard intermetallic parameter ∆EEuEu exp,gas, reflecting the adjacent interactions between two Eu(III) cations in the gas-phase complex, to compare it with its apparent small values found in solution ∆EEuEu exp,sol ) 14 kJ mol-1. Application of the site binding model32,33 to equilibria 14-16 in the gas phase gives eqs 25-27,33 from which the successive constants K2 (eq 28) and K3 (eq 29), corresponding to equilibria 17 and 18 in the gas phase, can be calculated (k is the specific affinity of Eu(III) for a nine-coordinate EuEu binding site and uEuEu ) e-∆Eexp,gas/RT is the Boltzmann factor containing the intermetallic interaction parameter ∆EEuEu exp,gas). βt13 ) 2k

(25)

βct23 ) 2k2uEuEu

(26)

3 EuEu 5/2 ) βtct 33 ) k (u

(27)

K2 ) βct23/βct13 ) kuEuEu

(28)

EuEu 3/2 ct ) K3 ) βtct 33/β23 ) (k/2)(u

(29)

The ultimate transformation of the successive stability constants into Gibbs free energy in eq 30 allows the -1 calculation of ∆EEuEu exp,gas ) 1500 kJ mol , a value in line with the predicted electrostatic interaction WEuEu elec,calcd ) 1388 kJ mol-1, operating between two triply charged Eu(III) cations separated by 9 Å in the gas phase (eq 4), but completely different from the apparent parameter of -1 found in solution. ∆EEuEu exp,sol ) 14 kJ mol 0 0 ∆GK3,gas - ∆GK2,gas ) - RTln(K3/K2) ) -

RTln(2) +

∆EEuEu exp,gas 2

( )

RT uEuEu ln ) 2 2

) 752 kJ mol-1 (30)

3520 Inorganic Chemistry, Vol. 46, No. 9, 2007

Experimental Section Solvents and Starting Materials. These were purchased from Fluka AG (Buchs, Switzerland) and Aldrich and were used without further purification unless otherwise stated. The ligand L5 was prepared according to a literature procedure.13 (nBu4N)4EDTA‚ 6.4H2O was prepared by metathesis of nBu4NOH and H4EDTA in water. The trifluoromethanesulfonate salts Ln(CF3SO3)3·xH2O (Ln ) La-Lu; x ) 1-4) were prepared from the corresponding oxides (Rhodia, 99.99%).34 The Ln content of the solid salts was determined by complexometric titrations with Titriplex III (Merck) in the presence of urotropine and xylene orange.35 Acetonitrile and dichloromethane were distilled over calcium hydride and nBu4NPF6 was recrystallized from hot ethanol. Preparation of RuCl2(DMSO)4. RuCl3 (600 mg, 2.9 mmol) was dissolved in DMSO (5 mL) and water (160 µL). The mixture was stirred and refluxed for 5 min, and then it was concentrated to half of its original volume. Acetone (50 mL) was added, and the resulting pale yellow solid was filtered, washed with diethyl ether, and dried overnight. The solid was recrystallized from DMSO (2 mL), filtered, washed with diethyl ether, and dried under vacuum to yield RuCl2(DMSO)4 (700 mg, 50%) as pale green-yellow crystals. Calcd for C8H24Cl2O4RuS4 (484.5 g mol-1): C, 19.83; H, 4.99. Found: C, 19.76; H, 4.96. Preparation of Ru(DMSO)6(CF3SO3)2‚H2O. RuCl2(DMSO)4 (720 mg, 1.49 mmol) and Ag(CF3SO3) (840 mg, 3.27 mmol, 2.2 equiv) were suspended in toluene (30 mL) and DMSO (420 µL, 5.94 mmol, 4 equiv). The light-protected mixture was refluxed for 1 h. The precipitate containing both AgCl and the insoluble Ru(II) salt was filtered, washed with toluene, and then extracted with methanol at 50 °C for 24 h. The resulting yellow solution was filtered, and the clear filtrate was evaporated to dryness to afford a pale yellow oil. The oil was dissolved in a mixture of acetone and diethyl ether and stored at room temperature. After 24 h, pale yellow crystals of Ru(DMSO)6(CF3SO3)2‚H2O (644 mg, 49%) were separated by filtration and dried under vacuum. Calcd for C14H36F6O12RuS8‚H2O (886.0 g mol-1): C, 18.97; H, 4.32. Found: C, 18.88; H, 4.30. Preparation of HHH-[RuLu(L5)3](CF3SO3)5‚3.5H2O (3). L5 (300 mg, 0.552 mmol, 3 equiv) and Lu(CF3SO3)3‚4.6H2O (130 mg, 0.184 mmol, 1 equiv) were suspended in ethanol (24 mL). The mixture was stirred at 50 °C for 30 min until complete dissolution of the solid. Ru(DMSO)6(CF3SO3)2‚H2O (163 mg, 0.184 mmol, 1 equiv) was added, and the mixture was stirred and refluxed for 24 h. The red mixture was stored at room temperature without stirring for 24 h. The red precipitate was separated by filtration and washed with ethanol. The resulting solution was evaporated to dryness to give a red solid. Recrystallization from CHCl3/ethanol afforded HHH-[RuLu(L5)3](CF3SO3)5‚3.5H2O (122 mg, 24%) as a red solid. Calcd for C104H99N21F15LuO18RuS5‚3.5H2O (2715.4 g mol-1): C, 46.00; H, 3.93, N, 10.83. Found: C, 45.94; H, 3.56; N, 10.65. ESIMS: m/z 514.4 ([RuLu(L5)3(CF3SO3)]4+). 1H NMR (CD3CN): δ 0.67 (t, 3J ) 7.1 Hz, 9H, Me5), 1.05 (t, 3J ) 7.1 Hz, 9H, Me4), 2.10 (s, 9H, Me1), 2.63 (m, 3H, H18), 2.79 (m, 3H, H16), 3.42 (m, 6H, H17,18), 3.61 (dd, 2J ) 16.4 Hz, 6H, H7,8), 4.35 (s, 9H, Me3), 4.43 (s, 9H, Me2), 4.87 (s, 3H, H6), 5.44 (s, 3H, H9), 6.99 (dd, 3J ) 8.4 Hz, 4J ) 1.4 Hz, 3H, H10), 7.22 (dd, 3J ) 8.5 Hz, 4J ) 1.3 Hz, 3H, H5), 7.35 (d, 4J ) 1.5 Hz, 3H, H1), 7.53 (d, 3J ) 8.4 Hz, (34) Desreux, J.-F. In Lanthanide Probes in Life, Chemical and Earth Sciences; Bu¨nzli, J.-C. G., Choppin, G. R., Eds.; Elsevier Publishing Co: Amsterdam, 1989; Chapter 2, p 43. (35) Schwarzenbach, G. Complexometric Titrations; Chapman & Hall: London, 1957; p 8.

Stabilities of Polymetallic Complexes 3H, H11), 7.63 (d, 3J ) 8.5 Hz, 3H, H14), 7.71 (dd, 3J ) 8.4 Hz, 4J ) 1.3 Hz, 3H, H4), 7.82 (d, 3J ) 8.0 Hz, 3H, H2), 8.28 (m, 6H, H3,13), 8.65 ppm (d, 3J ) 8.1 Hz, 3H, H3). Preparation of HHH-[Ru(L5)3](CF3SO3)2‚2.5H2O (1). A solution of 45 µmol of (nBu4N)4EDTA‚6.4H2O in acetonitrile (1.2 mL) was added dropwise to a solution of HHH-[RuLu(L5)3]CF3SO3)5‚ 3.5H2O (122 mg, 45 µmol) in acetonitrile (7.5 mL). The mixture was evaporated to dryness. The resulting red solid was dissolved in chloroform, and the pale red precipitate of (nBu4N)[LuEDTA] was removed by centrifugation. The remaining clear red solution was filtered and concentrated. Precipitation was induced by diffusion of diethyl ether. The precipitate was filtered off and washed with diethyl ether. This procedure was repeated three times to quantitatively remove the excess of nBu4N(CF3SO3). The final solid was dissolved in dichloromethane/tetrahydrofuran and then precipitated by the addition of diisopropyl ether. Filtration, followed by washing with diethyl ether and drying under vacuum at 60 °C, yields HHH-[RuL3](CF3SO3)2‚2.5H2O (76 mg, 81%) as an orange powder. Calcd for C101H99N21F6S2O9Ru‚2.5H2O (2075.24 g mol-1): C, 58.46; H, 5.05; N, 14.17. Found: C, 58.47; H, 5.04; N, 14.13. ESI-MS: m/z 866.8 ([Ru(L5)3]2+). 1H NMR (CD3CN): δ 1.05 (t, 3J ) 7.1 Hz, 9H, Me5), 1.23 (t, 3J ) 7.1 Hz, 9H, Me4), 2.18 (s, 9H, Me1), 3.31 (m, 6H, H17,18), 3.54 (m, 12H, H7,8,15,16), 4.11 (s, 9H, Me2), 4.13 (s, 9H, Me3), 5.71 (s, 3H, H6), 6.61 (dd, 3J ) 8.4 Hz, 4J ) 1.5 Hz, 3H, H10), 6.91 (s, 3H, H9), 7.20 (d, 3J ) 8.4 Hz, 3H, H11), 7.33 (dd, 3J ) 8.6 Hz, 4J ) 1.5 Hz, 3H, H5), 7.49 (s, 3H, H1), 7.53 (m, 6H, H4,14), 7.75 (dd, 3J ) 8.4 Hz, 4J ) 1.3 Hz, 3H, H2), 8.04 (t, 3J ) 8.0 Hz, 3H, H13), 8.27 (d, 3J ) 8.4 Hz, 3H, H3), 8.44 (dd, 3J ) 8.0 Hz, 4J ) 1.0 Hz, 3H, H12). Preparation of HHH-[RuCa(L5)3](CF3SO3)4‚7H2O (2). HHH[Ru(L5)3](CF3SO3)2‚2.5H2O (15 mg, 7.2 µmol) and Ca(CF3SO3)2 (2.44 mg, 7.2 µmol) were dissolved in dry acetonitrile (7 mL). The mixture was stirred at room temperature for 20 min and then evaporated to dryness. The resulting solid was dissolved in acetonitrile (2 mL), and diisopropylether (7 mL) was added to the solution. The orange powder was filtered, washed with diethyl ether, and dried under vacuum to yield HHH-[RuCa(L5)3](CF3SO3)4‚7H2O (13.3 mg, 74%) as an orange powder. Calcd for C103H99N21F12S4O15RuCa‚7H2O (2494.5 g mol-1): C, 49.59; H, 4.56; N, 11.79. Found: C, 49.54; H, 4.41; N, 11.66. ESI-MS: m/z 443.3 ([RuCa(L5)3]4+), 640.6 ([(RuCa(L5)3)(CF3SO3)]3+), 1035.4 ([(RuCa(L5)3)(CF3SO3)2]2+), 2219.3 ([(RuCa(L5)3)(CF3SO3)3]+). 1H NMR (CD3CN): δ 0.62 (t, 3J ) 5.7 Hz, 9H, Me5), 0.72 (t, 3J ) 5.8 Hz, 9H, Me4), 2.10 (s, 9H, Me1), 2.54-2.61 (m, 3H, H16), 2.77-2.83 (m, 3H, H15), 2.97-3.03 (m, 3H, H17), 3.14-3.20 (m, 3H, H18), 3.50 (dd, 2J ) 16.3 Hz, 6H, H7,8), 4.31 (s, 9H, Me3), 4.33 (s, 9H, Me2), 4.98 (s, 3H, H6), 5.89 (s, 3H, H9), 6.86 (dd, 3J ) 8.5 Hz, 4J ) 1.5 Hz, 3H, H10), 7.10 (dd, 3J ) 8.4 Hz, 4J ) 1.3 Hz, 3H, H5), 7.34 (s, 3H, H1), 7.36 (d, 3J ) 8.5 Hz, 3H, H11), 7.49 (d, 3J ) 7.5 Hz, 3H, H14), 7.53 (d, 3J ) 8.4 Hz, 3H, H4), 7.70 (dd, 3J ) 8.5 Hz, 4J ) 1.1 Hz, 3H, H2), 8.04 (t, 3J ) 8.0 Hz, 3H, H13), 8.25 (d, 3J ) 8.5 Hz, 3H, H3), 8.27 (d, 3J ) 8.2 Hz, 3H, H12). Spectroscopic and Analytical Measurements. Electronic spectra in the UV-vis region were recorded at 20 °C from solutions in CH3CN with a Perkin-Elmer Lamβda 900 spectrometer using quartz cells of 0.1 and 1 mm path lengths. Spectrophotometric titrations were performed with a J & M diode array spectrometer (Tidas series) connected to an external computer. In a typical experiment, 20 mL of [Ru(L5)3]2+ in CH3CN (10-4 mol‚dm-3) was titrated at 25 °C with a solution of Ln(CF3CO2)3‚xH2O (10-3 mol dm-3) in the same solvent under an inert atmosphere. After each addition of 0.10 mL, the absorbance was recorded using Hellma optrodes (optical path length 0.1 cm) immersed in the thermostated titration

vessel and connected to the spectrometer. Mathematical treatment of the spectrophotometric data was performed with factor analysis36 and with the SPECFIT program.30 1H NMR spectra were recorded at 25 °C on a Bruker Avance 400 MHz. Chemical shifts are given in parts per million with respect to TMS. Diffusion experiments were recorded at 400 MHz proton-Larmor frequency at room temperature. The sequence corresponds to Bruker pulse program ledbpgp2s37 using stimulated echo, bipolar gradients, and longitudinal eddy current delay as the z filter. The four 2 ms gradient pulses have sine-bell shapes and amplitudes ranging linearly from 2.5 to 50 G cm-1 in 16 steps. The diffusion delay was 100 ms, and the number of scans was 16. The processing was done using a line broadening of 5 Hz, and the diffusion rates were calculated using the Bruker processing package. Pneumatically assisted electrospray (ESI-MS) mass spectra were recorded from 10-4 mol dm-3 solutions on Finnigan SSQ7000 and Applied Biosystems: MDX SCIEX: API 150 EX instruments. Cyclic voltammograms were recorded using a BAS CV-50W potentiostat connected to a personal computer. A three-electrode system consisting in a stationary Pt disk working electrode, a Pt counter electrode, and a nonaqueous Ag/AgCl reference electrode was used. (nBu)4NPF6 (0.1 mol dm-3 in MeCN) served as an inert electrolyte. The reference potential (E0 ) -0.16 V vs SCE) was standardized against [Ru(bipy)3](ClO4)2 (bipy ) 2,2′-bipyridyl).14 The standard scan speed was 100 mV s-1, and voltammograms were analyzed according to established procedures.14 Elemental analyses were performed by Dr. H. Eder from the Microchemical Laboratory of the University of Geneva, Switzerland.

Conclusion The reorganization of solvent molecules around the cationic complexes occurring upon Ru(III)/Ru(II)-centered reduction in [RuLn(L5)3](3+z)+ or upon complexation of Ln3+ to [M(L5)3]m+ (M ) Cr, Ru) is mainly responsible for the apparent lack of considerable intermetallic electrostatic communication in solution and for the surprising stabilities of highly charged self-assembled polymetallic complexes in solution. However, Born-Haber cycles combined with reasonable semiquantitative approximations for estimation of the solvation energies for large cationic complexes restore standard electrostatic trends in the gas phase which can be captured by a simple Coulombic model (i.e., monopolemonopole intermetallic interactions). These results bring some insight into the interpretation of the intermetallic parameters ∆EMM sol fitted by thermodynamic models in solution, which can be now estimated thanks to simple electrostatic calculations in the gas phases “corrected” by solvation energies. It is thus not necessary to invoke either some breakdown of the dielectric continuum (i.e., local dielectric constants)9 or some drastic polarization of the metallic centers for the rationalization of intercomponent interactions occurring in solution. We are now in a position to eventually reconsider the debatable intermetallic interaction accompanying the successive fixation of Cu(I) in famous Lehn’s doublestranded helicates [Cun(L8)2]n+ (n ) 1-3), in which the monovalent cation are separated by 5.8 Å.1 The use of (36) Malinowski, E. R.; Howery, D. G. Factor Analysis in Chemistry; Wiley: New York, 1980. (37) Wu, D.; Chen, A.; Johnson, C. S. Jr. J. Magn. Reson. A 1995, 115, 260-264.

Inorganic Chemistry, Vol. 46, No. 9, 2007

3521

Canard and Piguet Coulomb eq 4 in the gas phase predicts ) 239 kJ 0 mol-1 (d ) 5.8 Å, r ) 1), which translates into ∆GK3,gas 0 -1 ∆GK2,gas ) 121 kJ mol upon application of eq 30 for the successive complexation processes given in equilibria 31 and 32, which strictly mirror those described for [Eun(L3)3]3n+ in equilibria 17 and 18 (n ) 1-3). ∆ECuCu gas

t-[Cu(L8)2]+ + Cu+ h ct-[Cu2(L8)2]2+ log(K2) ) log(βct22) - log(βt12) (31) ct-[Cu2(L8)2]2+ + Cu+ h [Cu3(L8)2]3+ log(K3) ) ct log(βtct 32) - log(β22) (32)

Solvation energies of ∆solvG°(t-[Cu(L8)2]+) ) -104 kJ mol-1, ∆solvG°(ct-[Cu2(L8)2]2+) ) -369 kJ mol-1, and ∆solvG°([Cu3(L8)2]3+) ) -747 kJ mol-1 can be easily computed with Born eq 6 assuming that (i) r ) 23 for acetonitrile/dichloromethane (1:1),2 (ii) r([Cu3(L8)2]3+) ) 8.0 Å in agreement with the experimental spherical hydrodynamic radius in acetonitrile,23 and (iii) the removal of each Cu(I) on going from [Cu3(L8)3]3+ to ct-[Cu2(L8)2]2+ and finally to t-[Cu(L8)2]+ shrinks the cationic radius by stepwise 10% increments, as applied previously for the related [Eu3(L3)3]9+ system (i.e., r([ct-Cu2(L8)2]2+) ) 7.2 Å and r([tCu(L8)2]+) ) 6.4 Å). The introduction of these data into eq 0 0 - ∆GK2,sol ) 8 kJ mol-1, which eventu24 gives ∆GK3,sol CuCu ally translates into ∆Esol ) 12 kJ mol-1 with eq 30. The

3522 Inorganic Chemistry, Vol. 46, No. 9, 2007

latter estimation is in good agreement with ∆ECuCu exp,sol ) 7 kJ mol-1 extracted from the analysis of the thermodynamic data collected in solution.5b Finally, it is worth noting that the same reasoning with a stepwise contraction of only 5% on going from r([Cu3(L8)2]3+) ) 8.0 Å to r(ct-[Cu2(L8)2]2+) ) 7.6 Å and r(t-[Cu(L8)2]+) ) 7.2 Å would eventually provide ) -43.4 kJ mol-1, which indeed corresponds to an ∆ECuCu sol apparent attraction between Cu(I) cations upon their successive fixation within the double-stranded mold in solution. It is thus not completely absurd to obtain negative intermetallic interaction parameters ∆EMM exp,sol < 0 when the multicomponent assemblies of charged species are analyzed with standard thermodynamic models. We are currently working on the concept of cooperativity associated with these intercomponent parameters because of the dependence on solvation energies. Acknowledgment. We thank Prof. M. Borkovec for fruitful discussions on solvation energies and Dr P.-Y. Morgantini for constructing Connolly surfaces. Financial support from the Swiss National Science Foundation is gratefully acknowledged. Supporting Information Available: Figures S1 shows the crystal structures of the cations [Ru(bipy)3]2+ and [RuLu(L5)3]5+ with Connolly surfaces. This material is available free of charge via the Internet at http://pubs.acs.org. IC062126O